<<

First-principles modeling of fractionation between chloride-bearing molecules and minerals Etienne Balan, Laura Créon, Chrystèle Sanloup, Jérôme Aléon, Marc Blanchard, Lorenzo Paulatto, Hélène Bureau

To cite this version:

Etienne Balan, Laura Créon, Chrystèle Sanloup, Jérôme Aléon, Marc Blanchard, et al.. First- principles modeling of chlorine isotope fractionation between chloride-bearing molecules and minerals. Chemical Geology, Elsevier, 2019, 525, pp.424-434. ￿10.1016/j.chemgeo.2019.07.032￿. ￿hal-02326016￿

HAL Id: hal-02326016 https://hal.archives-ouvertes.fr/hal-02326016 Submitted on 8 Nov 2020

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. 1

2

3 First-principles modeling of chlorine isotope fractionation between

4 chloride-bearing molecules and minerals

5

6

7 Etienne Balan1, Laura Créon1, Chrystele Sanloup1, Jérôme Aléon1, Marc Blanchard2,

8 Lorenzo Paulatto1, Hélène Bureau1

9

10

11

12 1Sorbonne Université, CNRS, IRD, MNHN, Institut de Minéralogie, de Physique des Matériaux et de 13 Cosmochimie (IMPMC), 4 place Jussieu, 75252 Paris cedex 05, France 14 2Géosciences Environnement Toulouse (GET), Observatoire Midi-Pyrénées, Université de Toulouse, CNRS, 15 IRD, UPS, 14 avenue E. Belin, 31400 Toulouse, France 16

17

18

19

20 Keywords : Cl , first-principles modeling, brines, Cl-bearing minerals, solar

21 nebula 22 Abstract

23

24 Equilibrium 37Cl/35Cl fractionation factors in selected molecules, Cl-bearing

25 crystalline solids, and silicates in which Cl occurs at trace or minor concentration level

26 are determined from first-principles calculations, within the density functional theory

27 (DFT) scheme. Results on benchmarking molecules and crystalline solids are consistent

28 with the previous theoretical study of Schauble et al. (2003). The present study further

29 documents the control of the isotopic fractionation properties of chlorine by its local

30 bonding environment. Chloromagnesite and chlorapatite display similar isotopic

31 fractionation properties due to relatively similar bonding environment. In contrast,

32 trace Cl in Mg-serpentine (lizardite) and Mg-amphibole (anthophyllite) are enriched in

33 37Cl with respect to chloromagnesite, due to the structural constraints exerted by the

34 host structure on the substituted ion. This effect is even more pronounced when Cl is

35 associated to hydroxylated cationic vacancies in forsterite. An effect of the local bonding

36 environment on the Cl isotopic fractionation properties is also inferred for Cl- ions in

37 saturated aqueous solutions. It explains the systematic departure between theoretical

38 and empirical reduced partition function ratio observed for the alkaline chlorides,

39 differing from the agreement observed for the hydrated Cl salts. The reduced partition

40 function ratio of Cl- ions in concentrated solution of alkaline chlorides is smaller from

41 that observed in dilute solutions by an amount potentially reaching 1‰ at 22°C. Finally,

42 the calculation of fractionation factors between gas (HCl(g), NaCl(g), KCl(g)) and solids

43 (sodalite, chlorapatite, halite, HCl trihydrate) which likely prevailed in the solar nebula,

37 44 sustains a model in which the Cl enrichment of HCl(g) is produced by a Rayleigh type

45 fractionation during chlorine condensation at temperatures between 400 and 500 K. 46 This model could explain the heavier isotopic composition observed for bulk Earth and

47 various chondrites compared to the nebular gas.

48 49 1. Introduction

50

51 The 37Cl/35Cl isotopic composition of chlorine in major Earth's reservoirs

52 (, evaporites, mantle) exhibits a limited range of values, with most of them

53 falling between -0.5 ‰ and +0.5 ‰ from the seawater composition (Eggenkamp, 2014;

54 Barnes and Sharp, 2017). Strong departures from these values are most often ascribed

55 to non-equilibrium kinetic processes affecting the Cl isotopic composition, which include

56 Cl loss during degassing processes (Sharp et al., 2010a,b) as well as diffusion and ion

57 filtration in porous media explaining negative values of sedimentary pore fluids (e.g.

58 Godon et al., 2004). At a larger scale, degassing processes are expected to explain the

59 heavier isotopic composition of planetary reservoirs, such as the Moon or the crust of

60 Mars, because of the preferential loss of light chlorine isotope (Sharp et al., 2010a;

61 Barnes and Sharp, 2017; Wang et al., 2019). Equilibrium isotopic fractionation of

62 chlorine in nature is generally expected to be of smaller magnitude because chlorine

63 mostly occurs under a single redox state, the reduced chloride ions, and the relative

64 mass difference between the isotopes is small. In addition, and although equilibrium

65 isotopic fractionation involving oxidized forms of chlorine such as perchlorates can be

66 large (Schauble et al., 2003), the photochemical formation of these unstable species is

67 not expected to fractionate the chlorine isotopes (Barnes and Sharp, 2017).

68 Few equilibrium fractionation factors of Cl isotopes have been experimentally

- - 69 determined. They include the HCl-Cl (aq), (Sharp et al., 2010b) and the Cl2-Cl (aq) (Giunta et

70 al., 2017) fractionation coefficients at near-ambient conditions; as well as fractionation

71 between chloride salts and saturated solutions (Eggenkamp et al., 1995, 2016). At high

72 temperature, Sharp et al. (2007) reported a -0.3 ‰ fractionation factor between

73 sodalite and NaCl(l) at 825 °C and Cisneros (2013) documented a +0.1 ‰ fractionation 74 factor between amphibole and NaCl-bearing solution at 700°C and 0.2 GPa. Observations

75 made on seafloor serpentinite and altered oceanic crust also point to a small positive

- - 76 serpentine-Cl (aq) and amphibole-Cl (aq) fractionation.

77 Theoretical isotopic fractionation factors between molecular species and

78 crystalline chlorides have been predicted by combining experimentally observed

79 vibrational frequencies and theoretically predicted frequency shifts due to the isotopic

80 substitution (Schauble et al., 2003). For molecules, theoretical predictions were based

81 on ab-initio quantum-mechanical calculations, whereas those for crystalline solids made

82 use of empirical force-fields. This set of theoretical fractionation coefficients revealed

83 significant equilibrium enrichment in 37Cl in oxidized chlorine species as well as in Cl-

84 bearing organic molecules. The fractionation between more reduced Cl2 and HCl

85 molecules was found to be consistent with experimental data and the role of the local

86 environment of chloride ions in minerals was underlined, suggesting enrichment of

87 silicates in 37Cl with respect to simple Na, K or Rb chlorides (Schauble et al., 2003). More

88 recently, Schauble and Sharp (2011) reported a -0.7 ‰ fractionation between sodalite

89 and HCl at 950 K and a -0.02 ‰ fractionation between sodalite and NaCl(c) at 1098 K.

90 They also predicted a +3 to +6 ‰ fractionation between crystalline HCl hydrate and

91 HCl(g) at 140-160 K, a fractionation potentially explaining the variability of Cl isotopic

92 composition observed in chondrites (Sharp et al., 2013; Gargano et al., 2017).

93 Previous studies have shown that theoretical fractionation factors computed

94 from first-principles within density functional theory (DFT) provide useful information

95 on isotopic systems that are difficult to address on a purely experimental basis because

96 they involve, e.g., uncommon isotopic effects (Schauble et al., 2006), slow chemical

97 reactions (Méheut et al., 2007), weakly fractionating isotopes (Blanchard et al., 2009;

98 Moynier et al., 2011; Blanchard et al., 2017), incorporation of minor or trace elements in 99 minerals (Rustad and Zarzycki, 2008; Balan et al., 2018) or high-temperature reactions

100 between solids and dilute gas (Javoy et al., 2012). In the present work, we apply this

101 approach to chlorine isotopes by providing a set of theoretical fractionation factors

102 between selected molecules (Table 1) and crystalline solids (Table 2) in which chlorine

103 occurs as a major or a trace element, all systems being treated at the same theoretical

104 level. This study mostly focuses on chloride, the lowest redox state of chlorine, which

105 corresponds to the most frequently observed in natural systems. The investigated

106 phases include cosmochemically important molecules, (HCl(g), NaCl(g) and KCl(g)), Cl-

107 bearing minerals (chlorapatite, sodalite, chloromagnesite and Cl salts precipitating from

108 aqueous solutions), as well as rock-forming silicates displaying Cl as a trace element

109 (lizardite, anthophyllite, forsterite).

110

111 2. Methods

112

113 2.1. Expression of isotopic fractionation factors

114

115 Assuming a system with two stable isotopes, the equilibrium isotopic

116 fractionation coefficient of an element Y between two phases a and b, referred to as

117 α(a,b,Y), is defined by the isotopic ratios:

118

119 α(a,b,Y)=(Y*/Y')a/(Y*/Y')b (1)

120

121 where Y* and Y' are the abundances of the two different isotopes. Accordingly, it

122 can be related to the equivalent coefficients, referred to as β-factors, obtained for each

123 phase and a reference system corresponding to an ideal gas of Y atoms: 124

125 103 ln α(a,b,Y) = 103 ln β(a,Y) -103 ln β(b,Y) (2)

126

127 The β(a,Y) factor is related to the reduced partition function ratio of the isotopomers

128 and harmonic expressions have been derived for molecular systems by Bigeleisen and

129 Mayer (1947). Using the high-temperature product (Redlich-Teller) rule, the harmonic

130 β-factor of a diatomic molecule is given by:

131

* −hν ( kT ) ν * e−hν (2kT ) 1− e 132 β(a,Y) = * (3) ν −hν ( kT ) e−hν (2kT ) 1− e 133

134 were ν and €ν* are the harmonic stretching frequencies of molecules differing by a single

135 isotopic substitution, k the Boltzmann constant, h the Planck constant, and T the

136 temperature. This expression is valid for molecules containing more than a single Y

137 atom provided that the isotopic mixing is ideal, a verified assumption in the case of

138 chlorine isotopes (Schauble et al., 2003).

139 For a crystal, harmonic β-factors can be calculated in a similar way using the

140 following expression:

141

1/(NqN ) * * −hν q,i ( kT ) ⎡ 3Nat −hν q,i (2kT ) ⎤ ν q,i e 1− e β(a,Y) = ⎢ * ⎥ 142 ∏∏ −hν q,i ( kT ) −hν q,i (2kT ) (4) ⎣⎢ i=1 {q} ν q,i 1− e e ⎦⎥

143

144 €where νq,i are the frequencies of the phonon with wavevector q and branch index i =

* 145 1,3Nat. Nat is the number of atoms in the unit cell, νq,i and ν q,i are the vibrational 146 frequencies in two isotopically different materials, N is the number of sites for the Y

147 atom in the unit cell (e.g., Méheut et al., 2007). In the product over the Nq q-vectors of Eq.

148 (4), the three translational modes at the center of the Brillouin zone with ν0,i = 0 are not

149 considered. The β-factors can thus be calculated for molecules and crystalline solids

150 using Eq. (3) and (4) and vibrational frequencies calculated from first-principles.

151 The calculation of the whole vibrational spectrum is sometimes prevented by the

152 too large size of the considered system. In this case, a useful approximation can be used

153 to obtain the β-factor from the restoring force constants Fi corresponding to the

154 displacement of the isotopic atom in three mutually perpendicular directions

155 (Bigeleisen and Mayer, 1947; Moynier et al., 2011; Ducher et al., 2017):

156

3 M ʹ − M h 2 157 β ≈1+ F M M 96 2k 2T 2 ∑ i ʹ π i=1 (5)

158 where M and M' are the masses of the two isotopes.

159 €

160 2.2 Computational details

161

162 Relevant vibrational properties, i.e. phonon frequencies (Eq. 3 and 4) or restoring

163 force constants (Eq. 5), were obtained within the density functional theory (DFT)

164 framework, using the generalized gradient approximation (GGA) to the exchange-

165 correlation functional as proposed by Perdew, Burke and Ernzerhof (PBE; Perdew et al.,

166 1996) and a plane-wave / pseudopotential scheme, as implemented in the PWscf and

167 PHonon codes from the Quantum-Espresso package (Baroni et al., 2001; Giannozzi et al.,

168 2009; http://www.quantum-espresso.org). The ionic cores were described using

169 optimized norm-conserving Vanderbilt (ONCV) pseudopotentials (Hamann, 2013; 170 Schlipf and Gigy, 2015) with 80 and 480 Ry cut-offs on the electronic wave functions and

171 charge density, respectively, ensuring that total energies are converged within 1

172 mRy/atom. Before the computation of vibrational properties, the relaxation of atomic

173 internal coordinates was performed until the residual forces were less than 10-4 Ry/au.

174 Initial guess of mineral structures were built using the experimental data (Table 2)

175 obtained from the American Mineralogist Crystal Structure Database (Downs and Hall-

176 Wallace, 2003). To compute the properties of isolated molecules (Table 1), a single

177 molecule was inserted in a cubic box with cell parameter a = 15.87 Å, a size large enough

178 to marginalize the interactions between periodic images of the molecule. In this case, the

179 electronic integration was performed by restricting the Brillouin-zone sampling to a

180 single k-point. For minerals, k-point grids are reported in Table 2 and both the cell-

181 parameters and atomic positions were relaxed under zero pressure. Depending on the

182 system size a finite grid of q-points (Eq. 4) or the force-constant approximation (Eq. 5)

183 were used to compute the β-factors (Table 2). In the finite grids of q-vectors (Eq. 4), the

184 center of the Brillouin zone (Γ point) was not included.

185

186 2.3 Structural models of Cl incorporation in silicates at trace or minor

187 concentration level

188

189 Chlorine usually occurs in rock-forming minerals as a trace or minor element.

190 Hydrous phyllosilicates and amphiboles are able to incorporate chlorine up to the wt. %

191 level. These significant concentrations are explained by the possibility to substitute

192 chloride ions for structural OH groups in these hydrous minerals (Volfinger et al., 1985).

193 Oceanic serpentines can contain up to 6200 ppm Cl (e.g. Scambeluri et al., 2004) and are

194 considered as key actors of Cl recycling in the mantle (Kendrick et al., 2011). 195 Accordingly, periodic models of Cl-bearing lizardite (Mg3Si2O5(OH)4) and anthophyllite

196 (Mg7Si8O22(OH)2) have been considered for their relative chemical simplicity. In order to

197 minimize, as much as possible, spurious interactions between the periodic images of the

198 substituted chloride ions, the unit-cell of the periodic Cl-bearing lizardite model has

199 been built from a 2x2x2 supercell containing 112 atoms. Two different lizardite models

200 were considered, referred to as Cl1 and Cl2, corresponding to Cl substitution for an

201 inner-OH or an inter-layer OH , respectively. The anthophyllite cell also contains

202 two non-equivalent OH-groups leading to two different models, referred to as Cl1 and

203 Cl2. The local environment around the two OH-groups is however more similar than in

204 lizardite. In anthophyllite, the primitive cell contains 164 atoms and it was not possible

205 to handle a larger supercell. The distance between Cl in neighboring cells is 5.27 Å along

206 the c direction. As the structure is constrained along this direction by the silicate chains,

207 we expect that the local structure around chlorine ions, and therefore their isotopic

208 fractionation properties, is weakly affected by the presence of the chlorine ions in

209 neighboring cells.

210 The incorporation mechanisms of in olivine are still debated, despite its

211 potentially important role in the storage of halogens in the upper mantle (Beyer et al.,

212 2012; Dalou et al., 2012; Joachim et al., 2015; 2017). Several studies have pointed out

213 the key role of hydroxylated defects in the incorporation of in forsterite, the Mg

214 olivine end-member (Crépisson et al., 2014; Joachim et al., 2015). Compared with

215 fluorine, partition coefficients of chlorine between olivine and melt are smaller, with a

216 lower dependence on the olivine water content (Joachim et al., 2017). Nonetheless, in

217 absence of definite conclusions concerning the incorporation mechanism of chlorine in

218 olivine, we considered that the Cl for OH substitution in hydroxylated defects of olivine

219 still is a viable incorporation mechanism. Other mechanisms involving, e.g. the charge 220 compensation of cationic vacancies by the association of several chloride ions in a single

221 defect, does not seem to be more favorable because of the low Cl concentration.

222 Accordingly, two types of models of hydroxylated defects, related to Si or Mg vacancies

223 in forsterite, were considered as potential candidates for the structural incorporation of

224 chloride ions. These models have been built from a 2x1x2 forsterite supercell and

225 previously validated by comparison with infrared spectroscopic observations (Balan et

226 al., 2011; 2017). In these models, only one Cl atom has been substituted for one of the

227 OH groups. In the case of the hydroxylated Si vacancy, three non-equivalent OH groups

228 can be substituted, leading to three Cl-bearing forsterite models referred to as Si_Cl1,

229 Si_Cl2 and Si_Cl3 (Table 3), where the number corresponds to the substituted hydroxyl

230 and the configuration is identical to that of the fluorinated models investigated by

231 Crépisson et al. (2014). An additional model (Si_Cl1b) was considered, corresponding to

232 a configuration in which the O2-H group points toward the neighbouring vacant site as

233 in the hydroxylated Si vacancy model described in Balan et al. (2017). Five models

234 preserving the electrostatic charge neutrality can be built for the hydroxylated Mg

235 vacancy (Table 3).

236 The equilibrium geometry of Cl-poor silicates was obtained by displacing all

237 atoms up to a minimum energy state, characterized by the fact that the forces

238 experienced by the atoms vanish. During this relaxation, the cell geometry was kept

239 fixed to that of the pure phase, as previously done in e.g. Balan et al. (2014) or Crépisson

240 et al. (2018).

241

242 3. Results

243

244 3.1 Benchmarking the modeling approach 245

246 The first-principles calculation of isotopic fractionation factors relies on a

247 number of approximations which can introduce systematic bias in the computed values

248 of β-factors. The stronger approximations rely on the use of the harmonic

249 approximation and in the choice of the exchange-correlation functional, here the PBE

250 functional. As observed in previous studies, the PBE calculations tend to overestimate

251 bond lengths and crystal cell-parameters by typically 1 to 2 % (Table 1 & 2). The

252 agreement is slightly poorer for chlorapatite, with a stronger overestimation of the a

253 cell-parameter and a slight contraction of the c cell-parameter (Table 1). Similar

254 distortion of the cell geometry was observed in previous modeling of the properties of F-

255 and OH-apatite at the GGA level (Corno et al., 2006; Balan et al., 2011b; Aufort et al.,

256 2018). An overestimation of the c cell-parameter of chloromagnesite is also observed. It

257 is likely related to the dominant Van der Waals contribution to the cohesion energy of

258 the layered chloromagnesite structure, a contribution which is not accounted for by the

259 PBE functional. Similar overestimation is observed in other layered minerals displaying

260 non -bonded neutral layers, such as talc. It is however not expected to strongly

261 affect the vibrational properties of the structure as the layer geometry is preserved

262 (Blanchard et al. 2018). Significant differences between theory and experiment are also

263 observed in the cell geometry of the hydrated salts bischofite (MgCl2.6(H20)) and

2+ 264 BaCl2.2(H20) (Table 2). The structure of bischofite displays Mg(OH2)6 polyhedra

- 265 sharing H-bonds with neighboring Cl ions or H2O molecules; whereas the hydrated

266 dichloride displays layers of edge sharing Ba(OH2)4Cl4 polyhedra. In such type of

267 molecular edifices, the neglect of Van der Waals interactions and thermal effects related

268 to the anharmonicity of weak H-bonds could be stronger than in crystals whose

269 cohesion is ensured by iono-covalent bonding. 270 The properties of small molecules (HCl, Cl2) and simple chloride salts (Table 1 &

271 2) can be compared to their experimental counterparts and to previous calculations by

272 Richet et al. (1977) and Schauble et al. (2003) (Fig. 1). As expected, the systematic

273 overestimation of bond distances and cell parameters due to the GGA, to an

274 underestimation of vibrational frequencies by ~5% (Fig. 2). The difference is however

275 larger for CsCl(c) than for the other halides, reaching 25% despite similar agreement

276 between theoretical and experimental cubic cell parameters (Table 2). A modeling of

277 CsCl using a different exchange-correlation functional (PBEsol; Perdew et al., 2008)

278 provided a smaller a cell-parameter (4.06 Å), increasing the zone-center TO frequency

279 to 93.5 cm-1 and the β-factor at 22°C by 0.43 ‰. For the molecules, a systematic ~0.5 ‰

280 decrease of β-factor at 0°C with respect to the previously calculated values is observed

281 (Fig. 1). The equilibrium fractionation factor between Cl2 and halite at 0°C is 4.6 ‰, very

282 close to the 4.75 ‰ value computed by Schauble et al. (2003). Similarly the halite-

283 sylvite fractionation at 0°C is 0.9 ‰ both in the present study and that of Schauble et al.

284 (2003), confirming that systematic bias tend to compensate when combining β-factors

285 computed at the same theoretical level.

286 As mentioned in the Method section, the whole vibrational spectrum is hardly

287 obtained for systems having a large number of atoms in their unit-cell. In this case, it can

288 be useful to use an approximate expression that relates the β-factor to the restoring

289 force constant when the relevant atom is displaced from its equilibrium position (Eq. 5).

290 This approximation of the harmonic expression of the reduced partition function ratio

291 assumes that the force-constants are isotope-independent. It is valid when the ratio of

292 the relevant vibrational frequencies to the temperature and the relative mass difference

293 between the isotopes are sufficiently small (Blanchard et al., 2017). Accordingly, the use

294 of the force-constant approximation was tested on the forsterite_Si_ Cl1 model. This 295 model displays a relatively high β-factor indicating that the relevant vibrational

296 frequencies are among the highest observed among the investigated systems. The

297 comparison between the β-factor obtained using the full expression (Eq. 4) and the

298 approximate one (Eq. 5) indicate that for a temperature above 450 K, the difference is

299 smaller than 0.1 ‰. As the force constant approximation is obtained from a finite

300 expansion of the full harmonic expression of the β-factor, the quality of the

301 approximation is better for systems displaying smaller vibrational frequencies and

302 related smaller β-factor. For example, the difference between the force constant

303 approximation and the full calculation for the NaCl and KCl molecules is smaller than 0.1

304 ‰ at 298 K.

305

306 3.2 Local environment and isotopic β-factor of chloride in condensed phases

307

308 The isotopic fractionation properties of chlorine reflect its local bonding

309 environment. For all the investigated solid phases, the temperature dependence of the

310 reduced partition function ratios is mostly linear and the corresponding slope

311 coefficients are given in Table 4. In monochloride salts, chlorine displays a coordination

312 varying from 8 in CsCl to 6 in the other chlorides with halite structure. The

313 corresponding reduced partition function ratio at 22°C (Table 5) display a systematic

314 variation as a function of the cation ionic radius, the larger cations leading to smaller β-

315 factors (Fig. 3). In the hydrated Mg, Ca and Sr dichloride salts, the chloride ion is bonded

316 to water molecules belonging to the coordination sphere of the cations (Table 3).

317 Consistently, these phases display similar β-factors, not correlated with the cation

318 radius (Fig. 3). In the peculiar HCl tri-hydrate structure, the HCl molecule is dissociated

+ 319 and the chloride ions are surrounded by Zündel ions (H5O2 ) (Lundgren and Olovsson, 320 1967). The corresponding β-factor is larger than that observed in the hydrated

321 dichloride salts in which chloride ions are surrounded by water molecules (Table 4).

322 When solely bound to divalent cations, chlorine displays a three-fold

323 coordination. This environment is observed in chloromagnesite and chlorapatite, as well

324 as in the silicates anthophyllite and lizardite. The β-factor of chloromagnesite (6.2 ‰ at

325 0°C) is consistent with the values reported for the isostructural compounds FeCl2 and

326 MnCl2 (6.7 and 6.2 ‰ at 0°C, respectively) by Schauble et al. (2003). Despite an

327 apparently similar environment and Mg-Cl distances close to 2.5 Å, the β-factor of Cl-

328 sites in lizardite and anthophyllite are significantly higher than that of chloromagnesite

329 (Table 4, Fig. 4). This suggests that the environment of Cl atoms is more constrained

330 when Cl is incorporated as a substituting element in silicates than when it occurs as a

331 major constituent of the mineral. As a matter of fact, the Mg-Cl-Mg angle decreases from

332 93 ° in chloromagnesite to about 80° in lizardite and anthophyllite. The effect of the host

333 structure on the vibrational and isotopic fractionation properties of Cl is even stronger

334 when considering the Cl sites in forsterite. In the case of the defects related to Si

335 vacancies, the Mg-Cl distances are smaller than in the chloride minerals (Table 3).

336 Neighboring OH groups can also point toward the Cl ion, suggesting a weak additional

337 bonding to H atoms. For the defects related to Mg vacancies, Cl ions are bound to Si

338 atoms and the largest β-factors are observed. Although it is not possible to directly

339 compare the energy of chemically different incorporation mechanisms, it can be noticed

340 that the less stable configurations of a given defect are observed for the environments

341 leading to the shortest Cl-cation distances (Table 3).

342 Finally, chloride ions in sodalite display a lower 4-fold coordination states with

343 bonding to Na atoms. Compared with halite, the lowering of the coordination state is

344 balanced by a shortening of the Cl-Na bond leading to a very similar β-factor for both 345 minerals. The calculated fractionation factor between HCl and sodalite (0.7 ‰ at 950 K)

346 is in excellent agreement with the value proposed by Schauble and Sharp (2011).

347

348 4. Discussion

349

350 4.1 Chlorine isotope fractionation during precipitation from saturated solutions

351

352 The precipitation of chloride salts from concentrated aqueous solutions is an

353 important natural process leading to the formation of evaporitic rocks. The equilibrium

354 isotopic fractionation factors of chlorine between aqueous solutions and precipitating

355 NaCl, KCl and hydrated MgCl2 has been experimentally determined by Eggenkamp et al.

356 (1995). The moderate values, varying between +0.26 ‰ and -0.08 ‰ for NaCl(c) and

357 KCl(c), respectively, were found to be roughly consistent with the theoretical values

358 provided by Schauble et al. (2003). Although debated (Luo et al., 2014, 2015;

359 Eggenkamp et al., 2015), the initially proposed experimental values have been more

360 recently confirmed by Eggenkamp et al. (2016). This series was also complemented by

361 the experimental determination of fractionation factors between aqueous solutions and

362 Li, Cs, Rb, Ca and Sr salts (Eggenkamp et al., 2016). The comparison of theoretical

363 results with experimental fractionation factors is not straightforward because theory

364 only provides β-factors within a given level of approximation. Although fractionation

365 factors and β-factors are simply related by Eq. (2), the theoretical prediction of

366 equilibrium isotopic fractionation between solids and solutions is hampered by

367 computational constraints which make it difficult to treat the liquid and solid phases at

368 the same approximation level (Blanchard et al., 2017). Using an empirical electrostatic

369 model of ion solvation by water molecules, Czarnacki and Halas (2012) have computed 370 the β-factor of the aqueous chloride ion. Combined with the β-factor of the Cl2 molecule

371 computed by Schauble et al. (2003), a good agreement was observed between the

372 theoretical values and the experimental fractionation factor between Cl2 vapor and

373 aqueous chloride ions determined by Giunta et al. (2017). This suggests that the β-factor

374 determined by Czarnacki and Halas (2012) can account for equilibrium fractionation

375 properties of chloride in real aqueous solutions. Thus, the β-factor of the salts can be

376 empirically assessed from Eq. (2), assuming that the reduced partition function ratio of

377 aqueous chloride calculated by Czarnacki and Halas (2012) (1000ln(β)=3.13 ‰ at 22

378 °C) matches that of saturated solutions and using the fractionation factors determined

379 by Eggenkamp et al. (2016). Similarly, the β-factor of the Cl2 molecule can be obtained

380 using the experimental data of Giunta et al. (2017). The comparison of these empirical β-

381 factors with the present theoretical ones reveal a high level of consistency for the Cl2

382 molecule and the hydrated Li, Mg, Ca and Sr salts, with a minor theoretical

383 underestimation of the β-factors most likely due to the systematic bias related to the

384 PBE functional (Fig. 5). A more pronounced departure is observed for cubic salts (NaCl,

385 KCl, RbCl, CsCl) and hydrated BaCl2, with the empirical values being systematically

386 larger than the theoretical ones. The observed difference is always above 0.5 ‰ and can

387 reach 1.5 ‰ for CsCl. In comparison, the difference between empirical and theoretical

388 β-factors for Cl2 molecules is only 0.14 ‰, supporting the combination of the aqueous

389 chloride β-factor computed by Czarnacki and Hallas (2012) with experimental

390 fractionation factors to infer empirical β-factors of solids and molecules. Experimental

391 measurements also provide isotopic composition values with typical errors smaller than

392 ±0.1 ‰ (Eggenkamp et al. 2016). Assuming that the experiments correspond to isotopic

393 equilibrium conditions, this suggests that the observed departure between empirical 394 and theoretical β-factors for these salts does not solely arise from theoretical or

395 experimental inaccuracies but rather reflects a variation in the β-factor of Cl- ions in

396 concentrated aqueous solutions. The weak experimental fractionation observed

397 between brines and crystalline salts, compared to the theoretical variation of the β-

398 factor of solid phases (Fig. 3), suggests that the β-factor of Cl- ions in concentrated

399 solutions follows the trend observed in the solids. According to the present results, the

400 β-factor of Cl- ions in saturated solutions could be smaller than that observed in dilute

401 solutions. The observed difference is larger than 0.5 ‰ and could reach 1 ‰ for CsCl

402 (Fig. 6), taking into account the ~0.43 ‰ difference related to the lower accuracy of the

403 PBE functional to model CsCl(c). Potential role of ion pairing in solution was previously

404 suggested by Schauble et al. (2003) and Eggenkamp et al. (2016). Significant contact ion-

405 pairing in concentrated CsCl and RbCl solutions has been recently observed by X-ray

406 absorption fine structure (XAFS) spectroscopy and X-ray diffraction (XRD) (Pham and

407 Fulton, 2016; 2018). In contrast, CaCl2 solutions are dominated by solvent-separated ion

408 pairs (Pham and Fulton, 2013; Friesen et al., 2019). Although deserving further

409 investigation, these observations consistently point to a stronger modification of the

410 local environment of chloride ions in concentrated aqueous solutions for the larger

411 alkaline cations with a smaller ionic potential.

412

413 4.2 Chlorine isotope fractionation in rock-forming silicates

414

415 Equilibrium fractionation of chlorine isotopes among silicate minerals is constrained by

416 few experimental measurements. Sharp et al. (2007) reported a -0.3 ‰ fractionation

417 factor between sodalite and NaCl(l) at 825 °C and Cisneros (2013) determined a +0.1 ‰

418 fractionation factor between amphibole and NaCl-bearing solution at 700°C and 0.2 GPa. 419 A fractionation of +0.33 ‰ between structurally bound and water-soluble Cl in seafloor

- 420 serpentinites is documented by Sharp and Barnes (2004). A positive amphibole-Cl (aq)

421 fractionation is also consistent with observations made on altered oceanic crust (Barnes

422 and Cisneros, 2012) and metasomatised meta-gabbro (Kusebauch et al., 2015). Based on

423 the similarity between the local environment of chlorine in serpentine and amphiboles

424 and that observed in dichloride salts, a 37Cl enrichment with respect to parent fluids is

425 consistent with the theoretical constraints (Schauble et al., 2003). The present results

426 provide twice as large β-factors of chlorine occurring as a trace or minor element in

427 silicates as those of dichloride salts. Assuming that the Cl β-factor in high temperature

428 fluids does not exceed the theoretical predictions for aqueous Cl (Czarnacki and Halas,

429 2012), they confirm the 37Cl enrichment of chloride bearing silicates with respect to the

430 parent fluids. The equilibrium isotopic fractionation between silicates and fluids

431 remains however weak, not exceeding 2 ‰ at 400 °C.

432 Concerning trace chlorine occurring as defects in forsterite, the usually high

433 formation temperature of olivine leads to very small isotopic fractionation. The

434 fractionation factor between forsterite and halite is smaller than 0.3 ‰ at 1300 °C.

435 Although very weak, measurement of differences in the isotopic composition of

436 experimentally grown samples of forsterite in presence of NaCl might help to

437 discriminate between inclusion of small halite particles and structural incorporation of

438 chlorine. We note however that lower formation temperature (<610 °C) of Fe-rich

439 olivine has been reported in some chondrites (Ganino and Libourel, 2017) potentially

440 leading to larger Cl isotope fractionation.

441 Overall, the weakness of equilibrium isotopic fractionation factors of chlorine at

442 high temperature cannot explain the variations observed in Middle Ocean Ridge basalts

443 (Bonifacie et al., 2008) or olivine-hosted melt inclusions from island arcs basalts 444 (Bouvier et al., 2019). It sustains the use of Cl isotopes to identify mixing processes

445 involving more fractionated and shallow Cl reservoirs such as seawater, altered oceanic

446 crust serpentinites or sediments.

447

448 4.3 Isotopic fractionation during Cl condensation in the solar nebula

449

450 The behavior of Cl and related isotopic fractionation during the formation of the

451 solar system is still an open question. It was initially thought that the homogeneity of

452 isotopic composition between bulk Earth, chondritic meteorites and lightest values of

453 Moon samples reflected a nebular composition close to 0 ‰ (Sharp et al., 2007). The

454 lack of isotopic fractionation between Earth and chondrites is explained by the hydrous

455 character of silicate melts, leading to a compensation between the positive fractionation

456 between HCl and Cl dissolved in melts and the preferential loss of light isotopes during

457 HCl degassing (Sharp et al., 2010a). In contrast, the elevated 37Cl/35Cl ratio measured on

458 Moon samples is explained by NaCl(g) loss from anhydrous melts.

459 However, the Cl isotopic composition of several ordinary chondrites (Sharp et al.,

460 2013), some shergottites having preserved the Mars mantle composition (Sharp et al.,

461 2016) and several meteorites (Gargano et al., 2017), suggests a lighter isotopic

462 composition of the nebular gas, between -3 and -5 ‰ (Sharp et al., 2016). Starting from

463 this composition, the high-temperature condensation of chlorine, corresponding to the

464 formation of sodalite at 950 K (Fegley and Lewis, 1980; Lodders, 2003), is not expected

465 to strongly fractionate the Cl isotopes. At this temperature NaCl molecules are the

466 dominant speciation of chlorine in the nebular gas (Fegley and Lewis, 1980). Based on

467 the present calculations, NaCl(g) displays almost no fractionation with sodalite, whereas

468 HCl(g) in equilibrium with NaCl(g) is only 0.7 ‰ heavier. Accordingly, the significantly 469 heavier composition observed on various chondrites, as well as in bulk Earth, cannot be

470 explained by high-temperature fractionation mechanisms. For this reason, Sharp et al.

471 (2016) proposed that these heavy compositions could reflect the low-temperature (<

472 140K) addition of Cl as HCl hydrates formed beyond the snow line. Theoretical

473 fractionation factors proposed by Schauble and Sharp (2011), confirmed by the present

474 results, indicate that HCl hydrates could be enriched in 37Cl by up to 3-4 ‰ at these

475 temperatures (Fig. 7). This interpretation is consistent with the suggestion of Zolotov

476 and Miromenko (2007) that aqueous alteration of chondrites at low-pH could reflect the

477 incorporation of HCl hydrates at temperatures as low as 160 to 140 K. However, the

478 postulated enrichment in 37Cl of HCl hydrates would only correspond to a minor fraction

479 of hydrates in equilibrium with a dominant fraction of HCl gas having escaped from

480 high-temperature condensation processes and still not fully condensed at low

481 temperature. Otherwise, mass balance implies that the isotopic composition of HCl

482 hydrates should tend to that of the nebular gas as their low-temperature condensation

483 proceeds. Such a low fraction of HCl condensed as hydrates is at odd with observations

484 of protostars (Kama et al., 2015) and comet 67P/Churyumov-Gerasimenko (Dhooghe et

485 al., 2017), indicating that up to 90% of HCl, the major Cl-bearing species in the

486 interstellar medium is frozen as ice onto the surfaces of dust grains.

487 The above described scenario is strongly dependent on the assumptions made on

488 the high-temperature condensation of chlorine. Considering the condensation model

489 proposed by Schaefer and Fegley (2010), a different scenario of the evolution of the Cl

490 isotopic composition in the solar nebula could be proposed. According to the

491 thermodynamical modeling of Schaefer and Fegley (2010) and more recently of Wood et

492 al. (2019), chlorine condensates at a lower temperature, starting by forming

493 chlorapatite at 470K by reaction of previously formed whitlockite or fluorapatite with 494 HCl(g) up to about 40% of total chlorine, and ending by the condensation of halite

495 directly from HCl gas at temperatures below 420K. Such lower condensation

496 temperature of chlorine is also consistent with the recent reevaluation of chlorine

497 during Earth accretion by Clay et al. (2017). New measurements of Cl concentration in

498 carbonaceous, enstatite and primitive ordinary chondrites reveals that a lower

499 condensation temperature of chlorine, close to 500 K, would be consistent with the

500 general trend observed for the other volatile elements. Based on the present results,

501 HCl(g), which is the dominant Cl-bearing constituent of the nebular gas at these

502 temperatures (Schaefer and Fegley, 2010), is enriched by 0.7 ‰ with respect to

503 chlorapatite at 470K and 1.7 ‰ with respect to halite at 420K (Fig. 8). At these

504 temperatures, it is reasonable to assume that the solid grains were isolated from the gas

505 and do not re-equilibrate with the gas, sustaining a Rayleigh fractionation model in the

506 absence of volatilization. In its simplest expression, this model describes the evolution of

37 35 507 the Cl/ Cl ratio R of the remaining HCl(g) as a function of the condensed fraction of

508 mineral phase f:

R 509 = f (α −1) (6) R°

510 where R° is the initial ratio of the nebular gas and α the fractionation factor between the € 511 gas and the condensed phase. While a significant degree of supercooling or HCl(g)

512 supersaturation during condensation would request the use of kinetic fractionation

513 factors, equilibrium fractionation factors can be used in the theoretical framework of

514 condensation at thermodynamical equilibrium. Using the calculated fractionation

515 factors, it appears that the condensation of halite can to a significant enrichment in

37 516 Cl of the residual HCl(g), which is fractionated by more than +3 ‰ with respect to the

517 starting nebular composition for a condensed fraction above 90 % (Fig. 9). At this point 518 40% of nebular chlorine is locked in chlorapatite and 50% in halite, while 10% is still

519 present as HCl(g). The fractionation reaches more than +5.5 ‰ for a condensed fraction

520 above 98 % (in this case halite now accounts for 58% of total nebular chlorine and only

521 less than 2% are still present as HCl(g)). Later interaction of chondrites with heavier HCl

522 at lower temperature, still under gas state or condensed as hydrates, could lead to both

523 their acidic alteration (Zolotov and Miromenko, 2007) and isotopic enrichment in 37Cl as

524 proposed by Sharp et al. (2016).

525 The present scenario relies on an efficient condensation of halite as the source of

526 isotopic fractionation between nebular gas and chondrites and is only based on

527 equilibrium fractionation processes. However, the origin of halite in chondrites remains

528 unclear and more complex processes are likely involved. As for chlorapatite and

529 sodalite, halite probably formed during brine circulations on the parent bodies

530 (Zolensky et al., 1999, Whitby et al., 2000, Bridges et al., 2004) but some halite crystals

531 and their fluid inclusions have H and O isotopic systematics indicating preservation of

532 pristine nebular volatiles (Yurimoto et al., 2014). While the low Cl abundance in

533 chondrites (less than 10% of the solar Cl based on Clay et al. 2017) does not favor

534 efficient condensation of halite, we note that the Cl isotopic compositions of halite in the

535 Zag ordinary chondrite (-2.8‰ ≤ δ37Cl ≤ -1.4‰, Bridges et al., 2004) and sodalite in

536 carbonaceous chondrites inclusions (-2.1‰ ≤ δ 37Cl ≤ -0.4‰, Sharp et al., 2007) are

537 well in line with the composition expected in our Rayleigh distillation model.

538

539 5- Conclusion

540

541 In the present work, a consistent set of theoretical 37Cl/35Cl β-factors has been

542 computed for selected minerals and molecules at the density functional theory level. The 543 β-factor is an intrinsic property of the phases, defined with respect to an ideal system.

544 They are not directly accessible from experiment but their computation makes it

545 possible to relate the isotopic fractionation properties to local chemical bonding and to

546 disentangle the contribution of individual phases in equilibrium isotopic fractionation

547 processes.

548 The present results reveal the important role of solution chemistry on the

549 isotopic fractionation of chlorine during the precipitation of chloride salts from aqueous

550 solutions. They also suggest that the variability of chlorine isotopic composition

551 observed in objects of the solar system could be inherited from Rayleigh fractionation

552 processes during the nebula condensation.

553 Accordingly, further works should aim at developing efficient modeling methods

554 able to treat large systems such as chemically complex aqueous solutions and their

555 interfaces with solids at the best approximation level. The present results could also

556 motivate experimental works on isotopic fractionation processes during the high-

557 temperature condensation of solids and further measurements of Cl isotopic

558 composition in mineral paragenesis of primitive meteorites.

559

560 References

561

562 Agron, P.A., Busing, W.R., 1985. dichloride hexahydrate, MgCl2*6H2O, by

563 neutron diffraction. Acta Cryst. C41, 8-10.

564 Agron, P.A., Busing, W.R., 1986. and dichloride hexahydrates by

565 neutron diffraction. Acta Cryst. C 42, 141–143 566 Aufort, J., Ségalen, L., Gervais, C., Paulatto, L., Blanchard, M., Balan, E., 2017. Site-specific

567 equilibrium isotopic fractionation of , and calcium in apatite.

568 Geochim. Cosmochim. Acta 219, 57-73.

569 Balan, E., Ingrin, J., Delattre, S., Kovacs, I., Blanchard, M., 2011a. Theoretical infrared

570 spectrum of OH-defects in forsterite. Eur. J. Mineral. 23, 285–292.

571 Balan, E., Delattre, S., Roche, D., Segalen, L., Morin, G., Guillaumet, M., Blanchard, M.,

572 Lazzeri, M., Brouder, C., Salje, E.K.H. 2011b. Line-broadening effects in the

573 powder infrared spectrum of apatite, Phys. Chem. Minerals 38, 111-122.

574 Balan, E., Blanchard, M., Pinilla, C., Lazzeri, M., 2014. First-principles modeling of sulfate

575 incorporation and 34S/32S isotopic fractionation in different calcium carbonates.

576 Chem. Geol. 374–375, 84-91

577 Balan, E., Blanchard, M., Lazzeri, M., Ingrin, J., 2017. Theoretical Raman spectrum and

578 anharmonicity of tetrahedral OH defects in hydrous forsterite. Eur. J. Mineral. 29,

579 201-212.

580 Balan, E., Noireaux, J., Mavromatis, V., Saldi, G.D., Montouillout, V., Blanchard, M.,

581 Pietrrucci, F., Gervais, C., Rustad, J.R., Schott, J., Gaillardet, J. 2018. Theoretical

582 isotopic fractionation between structural in carbonates and aqueous boric

583 acid and borate ion. Geochim. Cosmochim. Acta 222, 117-129.

584 Barnes, J.D., Cisneros, M., 2012. Mineralogical control on the chlorine isotope

585 composition of altered oceanic crust. Chem. Geol. 326−327, 51−60.

586 Barnes, J.D., Sharp, J.D., 2017. Chlorine isotope geochemistry. Reviews in Mineralogy &

587 Geochemistry 82, 345-378.

588 Baroni, S., de Gironcoli, S., Dal Corso, A., Giannozzi, P., 2001. Phonons and related crystal

589 properties from density-functional perturbation theory. Reviews of Modern

590 Physics 73, 515-561. 591 Beyer, C., Klemme, S., Wiedenbeck, M., Stracke, A., Vollmer, C., 2012. Fluorine in

592 nominally fluorine-free mantle minerals: experimental partitioning of F between

593 olivine, orthopyroxene and silicate melts with implications for magmatic

594 processes. Earth Planet. Sci. Lett. 337–338, 1–9.

595 Bigeleisen, J., Mayer, M.G., 1947. Calculation of equilibrium constants for isotopic

596 exchange reactions. J. Chem. Phys. 15, 261-267.

597 Blanchard, M., Poitrasson, F., Méheut, M., Lazzeri, M., Mauri, F., Balan E., 2009. Iron

598 isotope fractionation between pyrite (FeS2), hematite (Fe2O3) and siderite (FeCO3):

599 a first-principles density-functional theory study. Geochim. Cosmochim. Acta 73,

600 6565-6578.

601 Blanchard, M., Balan, E., Schauble, E., 2017. Equilibrium fractionation of non-traditional

602 isotopes: A molecular modeling perspective. Reviews in Mineralogy and

603 Geochemistry 82, 27-63.

604 Blanchard, M., Méheut, M., Delon, L., Poirier, M., Micoud, P., Le Roux, C., Martin, F., 2018.

605 Infrared spectroscopic study of the synthetic Mg-Ni talc series. Phys. Chem.

606 Minerals 45, 843-854.

607 Bonifacie, M., Busigny, V., Mevel, C., Philippot, P., Agrinier, P., Jendrzejewski, N.,

608 Scambellui, M., Javoy, M., 2008. Chlorine isotopic composition in seafloor

609 serpentinites and high-pressure metaperidotites. Insights into oceanic

610 serpentinization and subduction processes. Geochim. Cosmochim. Acta 72, 126–

611 139.

612 Bouvier, A., Manzini, M., Rose-Koga, E., Nichols, A.R.L. 2019. Tracing of Cl input into the

613 sub-arc mantle through the combined analysis of B, O and Cl isotopes in melt

614 inclusions. Earth Planet. Sci. Lett. 507, 30-39. 615 Bridges, J., Banks, D.A., Smith, M., Grady, M.M., 2004. Halite and stable chlorine isotopes

616 in the Zag H3-6 breccia. Meteor. Planet. Sci. 39, 657-666.

617 Cisneros, M. 2013. An experimental calibration of chlorine isotope fractionation

618 between amphibole and fluid at 700 °C and 0.2 GPa. M.S. Thesis, University of

619 Texas at Austin.

620 Clay, P.L., Burgess, R., Busemann, H., Ruzie-Hamilton, L., Joachim, B., Day, J.M.D.,

621 Ballentine, C.J., 2017. Halogens in chondritic meteorites and terrestrial accretion.

622 Nature 551, 614-618.

623 Corno, M., Busco, C., Civalleri, B., Ugliengo, P., 2006. Periodic ab initio study of structural

624 and vibrational features of hexagonal hydroxyapatite Ca10(PO4)6(OH)2. Phys.

625 Chem. Chem. Phys. 8, 2464–2472.

626 Crépisson, C., Blanchard, M., Bureau, H., Sanloup, C., Withers, A.C., Khodja, H., Surblé, K.,

627 Béneut, K., Leroy, C., Giura, P., Balan, E., 2014).Clumped fluoride-hydroxyl defects

628 in forsterite: Implications for the upper mantle. Earth Planet. Sci. Lett. 390, 287-

629 295.

630 Crépisson, C., Blanchard, M., Lazzeri, M., Balan, E., Sanloup, C., 2018. New constraints on

631 Xe incorporation mechanisms in olivine from first-principles calculations.

632 Geochim. Cosmochim. Acta 222, 146-155.

− 633 Czarnacki, M., Halas, S., 2012. Isotope fractionation in aqua-gas systems: Cl2–HCl–Cl ,

- 2- 634 Br2–HBr–Br and H2S–S . Isotopes in Environmental and Health Studies 45, 55−64.

635 Dalou, C., Koga, K.T., Shimizu, N., Boulon, J., Devidal, J.L., 2012. Experimental

636 determination of F and Cl partitioning between lherzolite and basaltic melt. Cont.

637 Mineral. Petro. 163, 591–609.

638 Dhooghe, F., De Keyser, J., Altwegg, K., Briois, C., Balsiger, H., Berthelier, J.-J., Calmonte, U.,

639 Cessateur, G., Combi, M.R., Equeter, E., Fiethe, B., Fray, N.,Fuselier, S., Gasc, S., 640 Gibbons, A., Gombosi, T., Gunell, H., Hässig, M., Hilchenbach, M., Le Roy, L.,

641 Maggiolo, R., Mall, U., Marty, B., Neefs, E., Rème, H., Rubin, M., Sémon, T., Tzou, C.-Y.,

642 Wurz, P., 2017. Halogens as tracers of protosolar nebula material in comet

643 67P/Churyumov-Gerasimenko. Monthly Notices of the Royal Astronomical Society

644 472, 1336-1345.

645 Downs, R.T., Hall-Wallace, M., 2003. The American Mineralogist Crystal Structure

646 Database. Amer. Mineral. 88, 247-250

647 Ducher, M., Blanchard, M., Balan, E., 2016. Equilibrium isotope fractionation in Zn-

648 bearing minerals from first-principles calculations. Chem. Geol. 443, 87-96.

649 Eggenkamp, H.G.M., Kreulen, R., Koster van Groos, A.F., 1995. Chloride stable isotope

650 fractionation in evaporites. Geochim. Cosmochim. Acta 59, 5169−5175.

651 Eggenkamp, H.G.M., 2014. The geochemistry of stable chlorine and isotopes.

652 Springer, Berlin

653 Eggenkamp, H.G.M., 2015b. Comment on “Stable isotope fractionation of chlorine during

654 the precipitation of single chloride minerals “by Luo, C.-g., Xiao, Y.-k., Wen, H.-j., Ma,

655 H.-z., Ma, Y.-q., Zhang, Y.-l., Zhang, Y.-x. And He, M.-y. [Applied Geochemistry 47

656 (2014) 141-149]. Appl. Geochem. 54, 111–116.

657 Eggenkamp, H.G.M., Bonifacie, M., Ader, M., Agrinier, P., 2016. Experimental

658 determination of stable chlorine and bromine isotope fractionation during

659 precipitation of salt from a saturated solution. Chem. Geol. 433, 46-56.

660 Fegley, B. Jr., Lewis, J.S., 1980. Volatile element chemistry in the solar nebula: Na, K, F, CI,

661 Br, and P. Icarus 41, 439-455.

662 Friesen, S., Hefter, G., Buchner, R., 2019. Cation hydration and ion pairing in aqueous

663 solutions of MgCl2 and CaCl2. J. Phys. Chem. B 123, 891-900. 664 Fujino, K., Sasaki, S., Takeuchi, Y., Sadanaga, R., 1981. X-ray determination of electron

665 distributions in forsterite, fayalite and tephroite. Acta Cryst. B 37, 513–518.

666 Ganino, C., Libourel, G., (2017) Reduced and unstratified crust in CV chondrite parent

667 body. Nature Comm. 8, 261.

668 Gargano; A.M., Sharp, Z.D., Taylor, L.A., 2017. Further constraining the chlorine isotope

669 composition of the solar nebula: main group iron meteorites. 80th Annual

670 meeting of the Meteoritical Society (LPI contrib. N°1987), abstract#6141.

671 Giannozzi, P., Baroni, S., Bonini, N., Calandra, M., Car, R., Cavazzoni, C., Ceresoli, D.,

672 Chiarotti, G.L., Cococcioni, M., Dabo, I., Dal Corso, A., de Gironcoli, S., Fabris, S.,

673 Fratesi, G., Gebauer, R., Gerstmann, U., Gougoussis, C., Kokalj, A., Lazzeri, M.,

674 Martin-Samos, L., Marzari, N., Mauri, F., Mazzarello, R., Paolini, S., Pasquarello, A.,

675 Paulatto, L., Sbraccia, C., Scandolo, S., Sclauzero, G., Seitsonen, A.P., Smogunov, A.,

676 Umari, P., Wentzcovitch, R.M., 2009. Quantum ESPRESSO: a modular and open-

677 source software project for quantum simulations of materials. Journal of Physics:

678 Condensed Matter 21, 395502.

679 Giunta, T., Labidi, J., Eggenkamp, H.G.M., 2017. Chlorine isotope fractionation between

- 680 chloride (Cl ) and dichlorine (Cl2). Geochim. Cosmochim. Acta 213, 375–382.

681 Godon, A., Jendrzejewski, N., Castrec-Rouelle, M., Dia, A., Pineau, F., Boulègue, J., Javoy,

682 M., 2004. Origin and evolution of fluids from mud volcanos in the Barbados

683 accretionary complex. Geochim. Cosmochim. Acta 68, 2153–2165.

684 Gregorkiewitz, M., Lebech, B., Mellini, M., Viti, C., 1996. Hydrogen positions and thermal

685 expansion in lizardite-1T from Elba: A low-temperature study using Rietveld

686 refinement of neutron diffraction data. Amer. Mineral. 81, 1111–1116.

687 Hamann, D.R., 2013. Optimized norm-conserving Vanderbilt pseudopotentials. Phys.

688 Rev. B 88, 085117. 689 Hassan, I., Antao, S.M., Parise, J.B., 2004. Sodalite: High-temperature structures obtained

690 from synchrotron radiation and Rietveld refinements. Amer. Mineral. 89 359-

691 364.

692 Hendricks, S., Jefferson, M., Mosley, V. 1932. The crystal structures of some natural and

693 synthetic apatite-like substances. Zeit. Kristall. 81, 352-369.

694 Hoering, T.C., Parker, P.L., 1961. The geochemistry of the stable isotopes of chlorine.

695 Geochim. Cosmochim. Acta 23, 186−199.

696 Hönnerscheid, A., Nuss, J., Mühle, C., Jansen, M., 2003. Die kristallstrukturen der

697 monohydrate von lithiumchlorid und lithiumbromid. Zeit. anorganische und

698 allgemeine Chem. 629, 312-316.

699 Javoy, M., Balan, E., Méheut, M., Blanchard, M., Lazzeri, M., 2012. First-principles

700 investigation of equilibrium isotopic fractionation of O- and Si-isotopes between

701 refractory solids and gases in the solar nebula. Earth Planet. Sci. Lett. 319/320,

702 118-127.

703 Joachim, B., Pawley, A., Lyon, I.C., Marquardt, K., Henkel, T., Clay, P.L., Ruzié, L., Burgess,

704 R., Ballentine, C.J., 2015. Experimental partitioning of F and Cl between olivine,

705 orthopyroxene and silicate melt at Earth’s mantle conditions. Chem. Geol. 416,

706 65–78.

707 Joachim, B., Stechern, A., Ludwig, T., Konzett, J., Pawley, A., Ruzie-Hamilton, L., Clay, P.L.,

708 Burgess, R., Ballentine, C.J., 2017. Effect of water on the fluorine and chlorine

709 partitioning behavior between olivine and silicate melt. Contrib. Mineral. Petrol.

710 172, 15.

711 Kama, M., Caux, E., Lopez-Sepulcre, A., Wakelam, V., Dominik, D., Ceccarelli, C., Lanza, M.,

712 Lique, F., Ochsendorf, B.B., Lis, D.C., Caballero, R.N., Tiellens, A.G.G.M., 2015. 713 Depletion of chlorine into HCl ice in a protostellar core: The CHESS spectral

714 survey of OMC-2 FIR 4. Astronomy Astrophysics 574, A107.

715 Kendrick, M.A., Scambelluri, M., Honda, M., Phillips, D., 2011. High abundances of noble

716 gas and chlorine delivered to the mantle by serpentinite subduction. Nature Geos.

717 4, 807–812.

718 Kusebauch, C., John, T., Whitehouse, M.J., Engvik, A.K. 2015. Apatite as probe for the

719 composition of metamorphic fluids (Bamble Sector, SE Norway). Contrib.

720 Mineral. Petrol. 170, 34.

721 Lodders, K., 2003. Solar system abundances and condensation temperatures of the

722 elements. Astrophysical J. 591, 1220–1247.

723 Lundgren, J.-O., Olovsson, I., 1967. Hydrogen bond studies. XV. The crystal structure of

724 hydrogen chloride dihydrate. Acta Cryst. 23, 966-971.

725 Luo, C.G., Xiao, Y.K., Wen, H.J., Ma, H.H., Ma, Y.Q., Zhang, Y.L., Zhang, Y.X., He, M.Y., 2014.

726 Stable isotope fractionation of chloride during the precipitation of single chloride

727 minerals. Appl. Geochem. 47, 141–149.

728 Luo, C.G., Xiao, Y.K., Wen, H.J., Ma, H.H., Ma, Y.Q., Zhang, Y.L., He, M.Y., 2015. Reply to the

729 comment on the paper “Stable isotope fractionation of chlorine during the

730 precipitation of single chloride minerals”. Appl. Geochem. 54, 117–118.

731 Méheut, M., Lazzeri, M., Balan, E., Mauri, F., 2007. Equilibrium isotopic fractionation

732 between kaolinite, quartz and water: prediction from first-principles density-

733 functional theory. Geochim. Cosmochim. Acta 71, 3170-3181.

734 Moynier, F., Yin, Q.-Z., Schauble, E., 2011. Isotopic evidence of Cr partitioning into Earth's

735 core. Science 331,1417-1420.

736 Padmanabhan, V., Busing, W., Levy, H., 1978. Barium chloride dihydrate by neutron

737 diffraction. Acta Cryst. B34, 2290-2292. 738 Perdew, J.P., Burke, K., Ernzerhof, M., 1996. Generalized Gradient Approximation Made

739 Simple. Phys. Rev. Lett. 77, 3865–3868.

740 Perdew, J.P., Ruzsinszky, A., Csonka, G.I., Vydrov, O.A., Scuseria, G.E., Constantin, L.A.,

741 Zhou, X., Burke, K., 2008. Restoring the density-gradient expansion for exchange

742 in solids and surfaces. Phys. Rev. Lett. 100, 136406.

743 Pham, V.-T., Fulton, J.L., 2013. Ion-pairing in aqueous CaCl2 and RbBr solutions:

744 Simultaneous structural refinement of XAFS and XRD data. J. Chem. Phys. 138,

745 044201.

746 Pham, V.-T., Fulton, J.L., 2016. High-resolution measurement of contact ion-pair

747 structures in aqueous RbCl solutions from the simultaneous corefinement of

748 their Rb and Cl K- edge XAFS and XRD spectra, J. Sol. Chem. 45, 1061–1070.

749 Pham, V.-T., Fulton, J. L., 2018. Contact ion-pair structure in concentrated cesium

750 chloride aqueous solutions: An extended X-ray absorption fine structure study. J.

751 Electron Spectro. and Related Phenom. 229, 20-25.

752 Reissland, J.A., 1973. The physics of phonons (John Wiley and sons LTD, London-New

753 York- Sydney-Toronto, 1973)

754 Richet, P., Bottinga, Y., Javoy M., 1977. A review of hydrogen, carbon, , oxygen,

755 and chlorine stable isotope fractionation among gaseous molecules. Annual

756 Reviews of Earth and Planetary Sciences 5, 65–110.

757 Rustad, J. R., Zarzycki, P., 2008. Calculation of site-specific carbon-isotope fractionation

758 in pedogenic oxide minerals. PNAS 105, 0297–10301.

759 Scambelluri, M., Müntener, O., Ottolini, L., Pettke, T.T., Vannucci, R., 2004. The fate of B, Cl

760 and Li in the subducted oceanic mantle and in the antigorite breakdown fluids.

761 Earth Planet. Sci. Lett. 222, 217–234. 762 Schauble, E.A., Rossman, G.R., Taylor, H.P., 2003. Theoretical estimates of equilibrium

763 chlorine-isotope fractionations. Geochim. Cosmochim. Acta 67, 3267-3281.

764 Schauble, E.A., Ghosh, P., Eiler, J.M., 2006. Preferential formation of 13C–18O bonds in

765 carbonate minerals, estimated using first-principles lattice dynamics. Geochim.

766 Cosmochim. Acta 70, 2510-2529.

767 Schauble, E.A., Sharp, Z.D., 2011. Modeling isotopic signatures of nebular chlorine

768 condensation. Goldschmidt Conference Abstracts, Mineral. Mag. 21, 1810.

769 Schaefer, L., Fegley, B., Jr., 2010. Cosmochemistry pp. 347–377, In Principles and

770 Perspectives in Cosmochemistry: Lecture Notes of the Kodai School on "Synthesis

771 of Elements" Eds. A. Goswami and B.E. Reddy, Springer.

772 Schlipf, M., Gygi, F. 2015. Optimization algorithm for the generation of ONCV

773 pseudopotentials. Computer Physics Communications 196, 36.

774 Shannon, R.D., 1976. Revised effective ionic radii and systematic studies of interatomic

775 distances in halides and chalcogenides", Acta Cryst. A32, 751-767.

776 Sharp, Z.D., Barnes, J.D., 2004. Water soluble chlorides in massive seafloor serpentinites:

777 a source of chloride in subduction zones. Earth Planet Sci Lett 226, 243−254.

778 Sharp, Z.D., Barnes, J.D., Brearley, A.J., Chaussidon, M., Fischer, T.P., Kamenetsky, V.S.,

779 2007. Chlorine isotope homogeneity of the mantle, crust and carbonaceous

780 chondrites. Nature 446, 1062−1065.

781 Sharp, Z.D., Shearer, C.K., McKeegan, K.D., Barnes, J.D., Wang, Y.Q., 2010a. The chlorine

782 isotope composition of the Moon and implications for an anhydrous mantle.

783 Science 329, 1050–1053.

784 Sharp, Z.D., Barnes, J.D., Fischer, T.P., Halick, M., 2010b. A laboratory determination of

785 chlorine isotope fractionation in acid systems and applications to volcanic

786 fumaroles. Geochim. Cosmochim. Acta 74, 264–273. 787 Sharp, Z.D., Mercer, J.A., Jones, R.H., Brearley, A.J., Selverstone, J., Bekker, A., Stachel, T.,

788 2013. The chlorine isotopic composition of chondrites and Earth. Geochim.

789 Cosmochim. Acta 107, 189-204

790 Sharp ZD, Williams J, Shearer CK, Jr., Agee CB, McKeegan KD (2016) The chlorine isotope

791 composition of Martian meteorites 2. Implications for the early solar system and

792 the formation of Mars. Meteor. Planet. Sci., 51, 2111-2126.

793 Sirdeshmukh, D.B., Sirdeshmukh, L., Subhadra, K.G., 2001. Alkali halides A handbook of

794 physical properties, Springer, Berlin.

795 Volfinger, M., Robert, J.-L., Vielzeuf, D., Neiva, A.M.R., 1985. Structural control of the

796 chlorine content of OH-bearing silicates (micas and amphiboles). Geochim.

797 Cosmochim. Acta 49, 37–48.

798 Wang, Y., Hsu, W., Guan, Y., 2019. An extremely heavy chlorine reservoir in the Moon:

799 Insights from the apatite in lunar meteorites. Scientific Rep. 9, 5727.

800 Whitby, J., Burgess, R., Turner, G., Gilmour, J., Bridges, J., 2000. Extinct 129I in halite from

801 a primitive meteorite: evidence for evaporite formation in the early solar system.

802 Science 288, 1819-1821.

803 Wood, B.J., Smythe, D., Harrison, T. 2019. The condensation temperatures of the

804 elements: a reappraisal. Amer. Mineral. in the press

805 Wyckoff, R.W.G, 1963. Crystal Structures 1 pp 239-444 Second edition. Interscience

806 Publishers, New York, New York

807 Yurimoto, H., Itoh, S., Zolensky, M., Kusakabe, M., Karen, A., Bodnar, R., 2014. Isotopic

808 compositions of asteroidal liquid water trapped in fluid inclusions of chondrites.

809 Geoch. J. 48, 549-560. 810 Zolensky, M.E., Bodnar, R.J., Gibson, E.K. Jr, Nyquist, L.E., Reese, Y., Shih, C.-Y., Wiesmann,

811 H., 1999. Asteroidal water within fluid inclusion-bearing halite in an H5

812 chondrite, Monahans. Science 285, 1377-1379.

813 Zolotov, M.Y., Mironenko, M.V., 2007 Hydrogen chloride as a source of acid fluids in

814 parent bodies of chondrites. Lun. Planet. Sci. Conf. 38, 2340.

815

816 817 Figure captions:

818

819 Figure 1: Reduced partition function ratio of HCl(g), Cl2(g), halite (NaCl(c)), sylvite (KCl(c))

820 and RbCl(c). Continuous lines: this study; dotted lines: Schauble et al. 2003; dashed lines:

821 Richet al. 1977. Inset: comparison of β-factors at 0°C.

822

823 Figure 2: Comparison of theoretical and experimental vibrational frequencies in Cl2

824 molecule and alkaline chlorides.

825

826 Figure 3: Theoretical β-factors at 22°C of anhydrous (full circles) and hydrated (empty

827 circles) chloride salts as a function of the cationic radius. Note the correlation observed

828 for the anhydrous salts and the weak variations of β-factors observed among the series

829 of hydrated salts. Ionic radii from Shannon (1976) for cations in 6-fold coordination,

830 excepted Cs and Ba (8-fold coordination).

831

832 Figure 4: Theoretical β-factors of minerals for temperatures above 450 K. Note the

833 higher values observed for substitutional Cl in silicates. The less stable forsterite models

834 (Si_Cl2, Mg_Cl2, Mg_Cl3 and Mg_Cl4) are not displayed.

835

836 Figure 5: Comparison of empirical and theoretical β-factors at 22°C. Full and empty

837 circles correspond to anhydrous and hydrated salts, respectively. Note the larger and

838 systematic discrepancy observed for the heavier alkaline chlorides. The size of the

839 symbols corresponds to an error bar of +/- 0.1 ‰. The error bar on the empirical β-

- 840 factor of Cl2 corresponds to that of the Cl2-Cl fractionation factor at 25°C reported by

841 Giunta et al. (2017). 842

843 Figure 6: Estimated theoretical β-factor of aqueous chloride ions, obtained by combining

844 the present theoretical β-factors of solids and the fractionation factors reported by

845 Eggenkamp et al. (2016), reported as a function of the ionic potential (Z/rion) of the

846 associated cation. The highest values, observed for Li and alkaline-earth (Sr, Ca, Mg)

847 counter-cations and averaging to 2.92, likely matches the theoretical β-factor of water

848 coordinated Cl- ions. Departure from this value is ascribed to the formation of contact

849 ion pairs with large alkaline cations. Estimated errors bars are +/- 0.2 ‰ combining the

850 precision of experimental measurements and theoretical values.

851

852 Figure 7: Reduced partition function ratio of HCl(g) and HCl trihydrate. Note the cross-

37 853 over at 205 K, leading to significant Cl enrichment of HCl(g) at temperatures lower than

854 140 K as previously reported by Schauble and Sharp (2011).

855

856 Figure 8: Isotopic fractionation factors between HCl(g) and other Cl-bearing molecules

857 and condensates from the nebular environment. Depending on the condensation

858 models, formation of sodalite at 950 K (Lodders 2003; Fegley and Lewis 1980) only

859 leads to a weak isotopic fractionation of HCl(g) with other phases; whereas a later

860 chlorine condensation as chlorapatite and halite (Schaefer and Fegley 2010) can lead to

37 861 a more significant Cl enrichment of HCl(g) at temperatures between 400 and 500K.

862 Note that the Cl speciation in the nebular gas is dominated by NaCl at 950 K and by HCl

863 at temperatures below 800 K.

864

865 Figure 9: Isotopic composition of remaining HCl(g) in a simple scenario of Rayleigh

866 fractionation during chlorine condensation between 500 and 400 K. The chlorapatite 867 accounts for 40% of chlorine condensation. For condensed fraction above 90 %, the

868 remaining HCl(g) is fractionated by more than +3 ‰ with respect to a starting nebular

869 composition of δ37Cl= -3 ‰.

870 871

872 Table 1: Structure and vibrational stretching frequencies of diatomic molecules. 873 Experimental harmonic frequencies (Richet et al. 1977) are indicated in parenthesis. 874 d (Å) theo. - exp. ω 35Cl (cm-1) theo. - exp. ω 37Cl (cm-1) theo. - exp. (%) (%) (%)

Cl2(g) 2.01 (1.98) +1.5 538.4 (559.7) -3.8 523.7 (544.4) -3.8 HCl(g) 1.29 (1.27) +1.6 2888.2(2990.9) -3.4 2886.0 (2988.7) -3.4 NaCl(g) 2.38 352.7 348.9 KCl(g) 2.67 268.7 264.8 875 876 877 878 Table 2: Structural and modeling parameters of the investigated crystalline phases. (c), 879 (h), (m) and (o) stand for cubic, hexagonal, monoclinic and orthorhombic crystal 880 systems, respectively. 881

model crystal k-point atoms/ cell parameters theo. - exp. ω TO (cm-1) q-point syst. grid cell (Å) (%) grid halite NaCl (c) 4x4x4 8 a = 5.69 (5.64)a +0.9 154 (164)a 4x4x4 sylvite KCl (c) 4x4x4 8 a = 6.38 (6.29)a +1.4 129 (142) a 4x4x4 RbCl (c) 4x4x4 8 a= 6.70 (6.59)a +1.6 102 (116.5) a 4x4x4 CsCl (c) 6x6x6 2 a= 4.21 (4.12)a +2.1 74 (99.5) a 6x6x6 b chloromagnesite MgCl2 (h) 4x4x2 9 a = 3.68 (3.60) +2.2 4x4x2 c = 19.62 (17.59) +11.5 c bischofite MgCl2.6(H2O) (m) 2x2x2 42 a = 9.08 (9.86) -7.9 1x1x1 b = 7.39 (7.11) +3.9 c = 6.62 (6.07) +9.0 β = 100.6° (93.7 °) d LiCl.H2O (o) 2x2x2 40 a = 7.69 (7.58) +1.4 1x1x1 b = 7.77 (7.68) +1.2 c = 7.64 (7.62) +0.3 e BaCl2.2(H2O) (m) 2x2x2 36 a = 6.85 (6.72) +1.9 1x1x1 b = 11.76 (10.91) +7.8 c = 6.99 (7.13) -1.9 β =90.06° (91.1°) f CaCl2.6(H2O) (h) 3x3x3 21 a= 7.88 (7.88) +0.0 2x2x2 c= 4.04 (3.95) +2.2 f SrCl2.6(H2O) (h) 3x3x3 21 a= 7.97 (7.96) +0.0 2x2x2 c= 4.21 (4.12) +2.2 g HCl tri-hydrate HCl.3(H2O) (m) 3x2x2 32 a = 4.07 (3.99) +2.0 3x2x2 b = 12.25 (12.05) +1.7 c = 6.75 (6.70) +0.7 β = 101.1° (100.6 °) h chlorapatite Ca5(PO4)3Cl (h) 2x2x2 42 a = 9.86 (9.52) +3.6 fca c = 6.73 (6.85) -1.7 j sodalite Na8Al6Si6O24Cl2 (c) 2x2x2 46 a = 8.96 (8.89) +0.8 fca i anthophyllite Mg7Si8O22(OH)2 (o) 1x1x2 164 a = 18.78 (18.50) +1.5 fca b = 18.14 (17.90) +1.3 c = 5.33 (5.27) +1.1 j lizardite Mg3Si2O5(OH)4 (h) 3x3x3 18 a = 5.37 (5.33) +0.8 fca c = 7.35 (7.25) +1.4 k forsterite Mg2SiO4 (o) 1x1x1 a = 9.60 (9.5) +1.1 fca (2x1x2 112 b = 10.30 (10.19) +1.1 supercell) c = 12.07 (11.96) +0.9 882 883 References: aSirdeshmukh et al. (2001), bWyckoff (1963), cAgron and Busing (1985), 884 dHönnerscheid et al. (2003), e Padmanabhan et al. (1978), fAgron and Busing (1986), gLundgren 885 and Olovsson (1967), hHendricks et al. (1932), iHassan et al. (2004), jGregorkiewitz et al. (1996), 886 kFujino et al. (1981) 887 Table 3: Cl coordination in crystals and relative energy of substitutional Cl-defects. 888 model Energy* Coord. distances (Å) model Coord. distances (Å) kJ/mol forsterite Si_Cl1 9.7 3 Mg, 3H Mg: 2.44, 2 x 2.40 H: 2 x 1.98, 2.0 halite 6 Na 2.84 Si_Cl1b 4.0 3 Mg, 2H Mg: 2 x 2.39, 2.48 H: 2 x 1.95 sylvite 6 K 3.19 Si_Cl2 44.3 3 Mg, 1H Mg: 2.29, 2 x 2.31 H: 2.08 RbCl 6 Rb 3.35 Si_Cl3 0 3 Mg, 1H Mg: 2.33, 2.39, 2.42 H: 2.0 CsCl 8 Cs 3.65 Mg_Cl1 0 2 Mg, 1Si Mg: 2.39, 2.40 Si: 2.17 chloromagnesite 3 Mg 2.53 Mg_Cl2 83.5 2 Mg, 1Si Mg: 2.43, 2.47 Si: 2.04 bischofite 4 H 2.13, 2.17, 2.21, 2.23 Mg_Cl3 94.9 2 Mg, 1Si Mg: 2.41, 2.50 Si: 2.05 HCl tri-hydrate 4 H 2.03, 2.08, 2.11, 2.15 Mg_Cl4 56.4 2 Mg, 1Si Mg: 2.40, 2.51 Si: 2.08 chlorapatite 3 Ca 2.79 Mg_Cl5 2.1 2 Mg, 1Si Mg: 2.37, 2.51 Si: 2.14 sodalite 4 Na 2.74 lizardite Cl1 0 3 Mg 2.53 LiCl. H2O 4 Li, 2H Li: 2.56, 2x2.67, 2.93 H: 2x2.20 Cl2 47.4 3 Mg 2 x 2.42, 2.46 BaCl2.2(H2O) 2 H 2.11, 2.18 2 H 2.12, 2.19 anthophyllite Cl1 10 3 Mg 2.47, 2 x 2.48 CaCl2.6(H2O) 6 H 3x2.19, 3x2.28 Cl2 0 3 Mg 2.45, 2 x 2.5

SrCl2.6(H2O) 6 H 3x2.21, 3x2.26 889 *For each type of defect, the energy is computed with respect to that of the most stable configuration. 890 891 892 893 894 Table 4: Polynomial fits of β-factor of crystals and molecules. 895 model a * 106/T2 model a * b * c* d* 106/T2 range range forsterite Si_Cl1 0.985 0-5 halite 0.253 0-13.4 Si_Cl1b 0.944 0-5 sylvite 0.185 0-13.4 Si_Cl2 1.124 0-5 RbCl 0.162 0-13.4 Si_Cl3 0.988 0-5 CsCl 0.130 0-13.4 Mg_Cl1 1.080 0-5 bischofite 0.243 0-13.4 Mg_Cl2 1.295 0-5 LiCl.H2O 0.272 0-13.4 Mg_Cl3 1.293 0-5 BaCl2.2(H2O) 0.263 0-13.4 Mg_Cl4 1.173 0-5 CaCl2.6(H2O) 0.263 0-13.4 Mg_Cl5 1.071 0-5 SrCl2.6(H2O) 0.254 0-13.4 lizardite Cl1 0.645 0-5 chloromagnesite 0.471 0-5 Cl2 0.740 0-5 HCl tri-hydrate 0.319 -1.993 10-3 3.951 10-5 -3.629 10-7 0-45 anthophyllite Cl1 0.740 0-5 NaCl(g) 0.226 0-5 Cl2 0.725 0-5 KCl(g) 0.175 0-5 -3 -5 -7 sodalite 0.234 0-5 Cl2(g) 0.6756 -6.557 10 8.684 10 -8.648 10 0-13.4 -3 -4 chlorapatite 0.418 0-5 HCl(g) 0.9415 -0.1164 9.165 10 -2.728 10 0-13.4 896 897 *The coefficients are defined by the relation: 103ln(β)=ax+bx2+cx3+dx4 where x=106/T2 and T is the 898 temperature in Kelvin. Note that quadratic, cubic and quartic coefficients have no physical meaning and 3 6 2 899 simply aim at reproducing the curvature of 10 ln(β) as a function of 10 /T for Cl2, HCl molecule and HCl 900 trihydrate. 901 902 903 904 905 Table 5: Theoretical β-factor of chloride salts and Cl2 molecule at 22 °C. 906 model 1000 ln(β) at 22°C halite 2.9 sylvite 2.1 RbCl 1.9 CsCl 1.5 bischofite 2.8 LiCl.H2O 3.1 BaCl2.2(H2O) 3.0 CaCl2.6(H2O) 3.0 SrCl2.6(H2O) 2.9 Cl2(g) 7.0 907 908 909 Figure 1: Reduced partition function ratio of HCl(g), Cl2(g), halite (NaCl(c)), sylvite (KCl(c))

910 and RbCl(c). Continuous lines: this study; dotted lines: Schauble et al. 2003; dashed lines: 911 Richet al. 1977. Inset: comparison of β-factors at 0°C. 912 913 914 200 °C 100 °C 0 °C

y= 0.48 + 0.99 x

8 Cl 8 2 (g)

4 ) (Schauble et al. 2003) ) (Schauble β ) β

1000 x ( ln 1000 T= 0 °C 0 0 4 8 HCl 1000 x ln (β) PBE (g) NaCl 4 (c) 1000 x ( ln 1000 KCl (c) RbCl (c)

0 0 4 8 12

106/T2 915 916 917 918 919 920 921 922 923 Figure 2: Comparison of theoretical and experimental vibrational frequencies in Cl2

924 molecule and alkaline chlorides.

925 926 927 600 Cl y = 1,0467x 2

400 ) exp. -1 (cm

ω 200 NaCl KCl CsCl RbCl

0 0 200 400 600 ω (cm-1) PBE 928 929 930 931 932 Figure 3: Theoretical β-factors at 22°C of anhydrous (full circles) and hydrated (empty

933 circles) chloride salts as a function of the cationic radius. Note the correlation observed

934 for the anhydrous salts and the weak variations of β-factors observed among the series

935 of hydrated salts. Ionic radii from Shannon (1976) for cations in 6-fold coordination,

936 excepted Cs and Ba (8-fold coordination).

937 938 939

LiCl.H O CaCl .6(H O) 2 2 2 BaCl .2(H O) 3 2 2 SrCl .6(H O) NaCl 2 2 MgCl .6(H O) 2 2 ) PBE β

KCl 2

1000 x ln( 1000 RbCl

CsCl

T = 22 °C 1 0.6 1 1.4 1.8

ionic radius (10-10 m) 940 941 942 943 Figure 4: Theoretical β-factors of minerals for temperatures above 450 K. Note the 944 higher values observed for substitutional Cl in silicates. The less stable forsterite models 945 (Si_Cl2, Mg_Cl2, Mg_Cl3 and Mg_Cl4) are not displayed. 946 947 900 °C 400 °C 200 °C 6

forsterite_Mg

forsterite_Si

) 4 β anthophyllite lizardite

1000 x ( ln 1000 chloromagnesite 2 chlorapatite halite sodalite sylvite

0 0 2 4 106/T2 948 949 950 951 952 953 954 955 956 957 958 959 Figure 5: Comparison of empirical and theoretical β-factors at 22°C. Full and empty

960 circles correspond to anhydrous and hydrated salts, respectively. Note the larger and

961 systematic discrepancy observed for the heavier alkaline chlorides. The size of the

962 symbols correspond to error bars of +/- 0.1 ‰. The error bar on the empirical β-factor

- 963 of Cl2 corresponds to that of the Cl2-Cl fractionation factor at 25°C reported by Giunta et

964 al. (2017).

965 966

Cl 7 2

1:1 5 ) empirical β

Na Ba K Sr 3 Cs Li Mg Ca 1000 x ln( 1000 Rb

1 1 3 5 7 1000 x ln(β) PBE 967 968 969 970 971 Figure 6: Estimated theoretical β-factor of aqueous chloride ions, obtained by combining

972 the present theoretical β-factors of solids and the fractionation factors reported by

973 Eggenkamp et al. (2016), reported as a function of the ionic potential (Z/rion) of the

974 associated cation. The highest values, observed for Li and alkaline-earth (Sr, Ca, Mg)

975 counter-cations and averaging to 2.92, likely matches the theoretical β-factor of water

976 coordinated Cl- ions. Departure from this value is ascribed to the formation of contact

977 ion pairs with large alkaline cations. Estimated errors bars are +/- 0.2 ‰ combining the

978 precision of experimental measurements and theoretical values.

979 980 981

Li 3 Ca 2.92 Sr Mg

) Ba

β Na

K Rb 2 1000 x ln( 1000

Cs

T = 22 °C 1 0 1 2 3 Z/r ion 982 983 984 985 Figure 7: Reduced partition function ratio of HCl(g) and HCl trihydrate. Note the cross-

37 986 over at 205 K, leading to significant Cl enrichment of HCl trihydrate at temperatures

987 lower than 140 K as previously reported by Schauble and Sharp (2011).

988 989

990 991 992 993 994 995 996 997 998 999 Figure 8: Isotopic fractionation factors between HCl(g) and other Cl-bearing molecules

1000 and condensates from the nebular environment. Depending on the condensation

1001 models, formation of sodalite at 950 K (Lodders 2003; Fegley and Lewis 1980) only

1002 leads to a weak isotopic fractionation of HCl(g) with other phases; whereas a later

1003 chlorine condensation as chlorapatite and halite (Fegley and Schaefer 2010) can lead to

37 1004 a more significant Cl enrichment of HCl(g) at temperatures between 400 and 500K.

1005 Note that the Cl speciation in the nebular gas is dominated by NaCl at 950 K and by HCl

1006 at temperatures below 800 K.

1007 1008 1009

1010 1011 Figure 9: Isotopic composition of remaining HCl(g) in a simple scenario of Rayleigh

1012 fractionation during chlorine condensation between 500 and 400 K. The chlorapatite

1013 accounts for 40% of chlorine condensation. For condensed fraction above 90 %, the

1014 remaining HCl(g) is fractionated by more than +3 ‰ with respect to a starting nebular

1015 composition of δ37Cl= -3 ‰.

1016

1017 1018

4

halite (420 K) Cl (per mil) (per Cl 0 chlorapatite 37

δ (470 K)

-4 0 40 80 % condensation 1019